Stereocontrolled Preparation of Spirocyclic Ethers by Intramolecular

David J. Holt, William D. Barker, and Paul R. Jenkins , Jagannath Panda and Subrata Ghosh. The Journal of Organic Chemistry 2000 65 (2), 482-493...
0 downloads 0 Views 403KB Size
7860

J. Org. Chem. 1996, 61, 7860-7866

Stereocontrolled Preparation of Spirocyclic Ethers by Intramolecular Trapping of Oxonium Ions with Allylsilanes Leo A. Paquette* and Jinsung Tae Evans Chemical Laboratories, The Ohio State University, Columbus, Ohio 43210 Received July 10, 1996X

The stereoselectivity of the spontaneous intramolecular cyclization of 2-(benzenesulfonyl)-2-(4((trimethylsilyl)methyl)-4-pentenyl)tetrahydropyrans substituted by alkyl groups at various ring positions has been examined. For the 4- and 6-methyl derivatives, formation of the spirocyclic center occurs exclusively anti to the methyl. The outcome in the 5-methyl example is a 3.7:1 syn/ anti split. For the trans-4,6-dimethyl derivative, the substituents act in a reinforcing manner and direct cyclization uniquely in one direction. Both the cis and trans bicyclic ethers ring close on that π-surface of the intermediate oxonium ion syn to the angular hydrogen. The results are rationalized in terms of the predilection of the associated oxonium ions for nucleophilic capture via a chairlike or twist-boat transition state. Oxonium ion intermediates are playing an everincreasing role as important synthetic intermediates,1,2 and the stereochemistry of their nucleophilic capture is of considerable interest.3,4 Various methods have been developed for trapping the oxonium ion intramolecularly as a useful means for elaborating functionalized rings. Marko’s silyl-modified Sakurai reaction,5 Ley’s R-alkoxy sulfone approach to methylenecyclohexane systems,6 and the Marko-Krief spiro acetal synthesis7 all rely on the nucleophilic responsiveness of the allylsilane functionality toward transient oxonium ion centers. This wellmatched reactivity results in efficient C-C bond construction. Earlier work from this laboratory has focused on the fate of oxonium ions substituted at CR with a carbinol carbon.8 The conversion of 1 to 2 was found to initiate a pinacol-like Wagner-Meerwein rearrangement. The

steric and stereoelectronic factors involved in this isomerization have been investigated,9 and the products have served as useful mechanistic probes in numerous contexts.10 Several direct applications of this chemistry to total synthesis have also been achieved.11,12 A more expansive scrutiny of these reactive intermediates is undoubtedly warranted. A particularly informaX Abstract published in Advance ACS Abstracts, October 15, 1996. (1) Perst, H. Oxonium Ions in Organic Chemistry; Verlag Chemie/ Academic Press: Weinheim/Bergstr., Germany, 1971. (2) Deslongchamps, P. Stereoelectronic Effects in Organic Chemistry; Pergamon Press: Oxford, England, 1983. (3) Hopkins, M. H.; Overman, L. E.; Rishton, G. M. J. Am. Chem. Soc. 1991, 113, 5354. (4) (a) Pothier, N.; Goldstein, S.; Deslongchamps, P. Helv. Chim. Acta 1992, 75, 604. (b) Deslongchamps, P.; Dory, Y. L.; Li, S. Can. J. Chem. 1994, 72, 2021. (c) Deslongchamps, P.; Dory, Y. L.; Li, S. Helv. Chim. Acta 1996, 79, 41. (5) Marko, I. E.; Bayston, D. J. Tetrahedron 1994, 50, 7141. (6) Ley, S. V.; Kouklovsky, C. Tetrahedron 1994, 50, 835. (7) Marko, I. E.; Bailey, M.; Murphy, F.; Declercq, J.-P.; Tinant, B.; Feneau-Dupont, J.; Krief, A.; Dumont, W. Synlett 1995, 123. (8) (a) Paquette, L. A.; Lawhorn, D. E.; Teleha, C. A. Heterocycles 1990, 30, 765. (b) Negri, J. T.; Rogers, R. D.; Paquette, L. A. J. Am. Chem. Soc. 1991, 113, 5073. (c) Paquette, L. A.; Negri, J. T.; Rogers, R. D. J. Org. Chem. 1992, 57, 3947. (9) (a) Paquette, L. A.; Andrews, J. F. P.; Vanucci, C.; Lawhorn, D. E.; Negri, J. T.; Rogers, R. D. J. Org. Chem. 1992, 57, 3956. (b) Paquette, L. A.; Lanter, J. C.; Johnston, J. N. Submitted for publication.

S0022-3263(96)01306-0 CCC: $12.00

Scheme 1

tive case study involves the stereoselective synthesis of oxabicyclics that are typified by 8 (Scheme 1). The expectation was that protonation of 4 would trigger the formation of 5 and induce ring closure along two possible pathways. Rapid equilibration between two conformations of 5, depicted as 6 and 7, can be anticipated. Antiperiplanar intramolecular attack by the flanking allylsilane residue13 should follow. In the case of 6, direct conversion to the chairlike product 8 would materialize. Cyclization via 7 would lead first to the twist-boat conformer 94a and subsequently, for thermodynamic reasons, to the chair-chair geometry. (10) (a) Paquette, L. A.; Branan, B. M.; Friedrich, D.; Edmondson, S. D.; Rogers, R. D. J. Am. Chem. Soc. 1994, 116, 506. (b) Branan, B. M.; Paquette, L. A. J. Am. Chem. Soc. 1994, 116, 7658. (c) Paquette, L. A.; Branan, B. M. Heterocycles 1995, 40, 101. (d) Paquette, L. A.; Branan, B. M.; Rogers, R. D. J. Org. Chem. 1995, 60, 1852. (e) Paquette, L. A.; Branan, B. M.; Rogers, R. D.; Bond, A. H.; Lange, H.; Gleiter, R. J. Am. Chem. Soc. 1995, 117, 5992. (f) Paquette, L. A.; Branan, B. M.; Stepanian, M. Tetrahedron Lett. 1996, 37, 1721. (g) Paquette, L. A.; Stepanian, M.; Branan, B. M.; Edmondson, S. D.; Bauer, C. B.; Rogers, R. D. J. Am. Chem. Soc. 1996, 118, 4504. (11) (a) Paquette, L. A.; Lord, M. D.; Negri, J. T. Tetrahedron Lett. 1993, 34, 5693. (b) Paquette, L. A.; Dullweber, U.; Cowgill, L. D. Tetrahedron Lett. 1993, 34, 8019. (c) Lord, M. D.; Negri, J. T.; Paquette, L. A. J. Org. Chem. 1995, 60, 191. (d) Paquette, L. A.; Wang, H.-L. Tetrahedron Lett. 1995, 36, 6005. (e) Paquette, L. A.; Wang, H.-L. J. Org. Chem. 1996, 61, 5352. (12) Paquette, L. A.; Kinney, M. J.; Dullweber, U. Submitted for publication. (13) Fleming, I.; Dunogue`s, J.; Smithers, R. Org. React. 1989, 37, 57.

© 1996 American Chemical Society

Stereocontrolled Preparation of Spirocyclic Ethers Scheme 2

J. Org. Chem., Vol. 61, No. 22, 1996 7861 Table 1. Spontaneous Cyclization Reactions of Alkyl-Substituted 2-(Benzenesulfonyl)pyrans following Alkylation with 10

In the absence of additional stereochemical markers, these two pathways cannot be distinguished. The goals of the present undertaking were initially to determine feasibility in the parent system and subsequently to probe possible stereoelectronic bias by alkyl substitution resident at various positions on the pyran ring. Results The systematic investigation was initiated by metalation of dihydropyran with tert-butyllithium and alkylation with the known iodide 10,5 giving the sensitive product 11. Exposure of 11 to Dowex-50 in CH2Cl2 for 5 h led efficiently to first-formed 12 and its postequilibrated isomers 13 and 14 (Scheme 2). In an alternative approach, sulfone 1514 was deprotonated and condensed with 10 at -78 °C. Slow warming of the reaction mixture to room temperature resulted in spontaneous generation of the oxonium ion14 and direct generation of 12-14. Since the overall yields of the two synthetic pathways are rather comparable, further work has relied on the sulfone approach because of its convenience and the readiness with which the starting substituted pyrans can be prepared. Also, the conditions do not require the coaddition of a strongly acidic catalyst such as Dowex-50. Spirocycle 12 is notably sensitive to acid and experiences a prototropic shift readily during chromatography, while standing in CDCl3 solution, and the like. The varied product distributions recorded in Scheme 2 reflect the different reaction conditions and times and are intended to serve only as a reflection of the acid lability of 12. The structural assignments to the olefinic isomers 13 and 14 are based reliably on 2D-COSY experiments. The distinctive nature of their vinyl proton chemical shifts in CDCl3 (δ 5.35 for 13 and δ 5.54 for 14) easily distinguishes these isomers from each other as well as from 12 (δ 4.71, 4.66). The known sulfones 1615 and 1916 were readily procurable in reasonable amounts as diastereomeric mixtures, which were used without separation. Alkylation of 16 with 10 led directly to a 98:2 mixture of 17 and 18 (Table 1). When 19 was subjected to the same protocol, a clear colorless oil consisting of 20 and 21 in an 81:19 ratio was (14) Ley, S. V.; Lygo, B.; Sternfeld, F.; Wonnacott, A. Tetrahedron 1986, 42, 4333. (15) Boons, G.-J.; Entwistle, D. A.; Ley, S. V.; Woods, M. Tetrahedron Lett. 1993, 34, 5649. (16) Genicot, C.; Ley, S. V. Synthesis 1994, 1275.

produced. As in the case of 15, the yields were modest because of the high volatility of the resulting spirocyclic ethers and the associated losses incurred during workup. The stereoselectivities of the two ring closures were exclusive and in the same direction. NOE enhancements observed for 17 and 20, as revealed in A and B, confirm the proximity of the neopentyl allylic protons to the axial R-oxy protons shown. The ready isomerization of 17 to 18 and of 20 to 21 under acidic conditions demand that they be stereochemically related at the two stereogenic centers.

The 5-methyl substituted tetrahydropyranyl sulfone 22 was prepared according to Scheme 3. Conversion of

7862 J. Org. Chem., Vol. 61, No. 22, 1996 Scheme 3

acrolein to the bromine-substituted diallyl acetal 2817 set the stage for the free-radical cyclization18 that involved heating with tributylstannane in benzene under AIBN catalysis. The direct conversion of 29 to 22 proceeded with good efficiency. A notable aspect of the alkylative spirocyclization of 22 is formation of the two diastereomeric ethers 23 and 24 (3.7:1, Table 1). Although it did not prove possible to effect isomer separation, the 300 MHz 1H NMR spectrum of this mixture was sufficiently well-defined to allow acquisition of the NOE data depicted in C and D. Therefore, in contrast to 16 and 19, the major pathway followed by 22 is to generate the new C-C bond syn to the methyl substituent. In order to probe whether the 4- or 6-methyl group is more controlling of the stereochemistry, attention was next directed to the trans-4,6-dimethyl example 25. The starting point for this phase of the undertaking was the previously described lactone 30.19 Hydride reduction of 30 to diol 31 was followed by protection of the primary hydroxyl as the tert-butyldimethylsilyl derivative (Scheme 4). Subsequent Mitsunobu inversion20 of the secondary

Paquette and Tae

carbinol with p-nitrobenzoic acid21 gave 33 from which the doubly protected 35 and the mono alcohol 36 were conventionally crafted. The final two steps consisted of oxidation to the aldehyde level with pyridinium chlorochromate in the presence of 4 Å molecular sieves and direct treatment of this penultimate intermediate with benzenesulfinic acid in the presence of CaCl2. The assumption that complete inversion of configuration has materialized during the conversion of 32 into 33 was confirmed within experimental uncertainty by the highyield conversion of 25 into 26 and 27 (7.4:1, Table 1). As before, the isomer ratio and structural assignments were established using 1H NMR and particularly NOE methods directly on the product mixture. The proximity of an axial proton and an axial methyl group to the neopentyl allylic methylene proton pair proved particularly diagnostic, as summarized in E and F. The postisomerization of 26 to 27 parallels in its regioselectivity the formation of 18 and 21 and is taken to be a reflection of the greater thermodynamic stability of that internal olefin isomer carrying its double bond away from the pyran ring.

Scheme 4

In order to assess the role of conformational rigidity during these cyclizations, the rigid trans-fused bicyclic sulfone 40 and its conformationally flexible cis isomer 47 were targeted for study. In the first instance, the efficiency with which cyclohexene oxide undergoes ring cleavage in the presence of 3-butenylmagnesium bromide and copper(I) iodide22 was taken advantage of (Scheme 5). The formation of 37 in this way allowed for the subsequent ozonolysis in methanol. Following the workup with methyl sulfide, a 1:3 mixture of cyclic lactol 38 and acetal 39 could be obtained and transformed directly into 40. Low-temperature alkylation of the anion of 40 (17) Hsung, R. P. Synth. Commun. 1990, 20, 1175. (18) Stork, G.; Mook, R., Jr.; Biller, S. A.; Rychnovsky, S. D. J. Am. Chem. Soc. 1983, 105, 3741. (19) McKelvey, R. D.; Kawada, Y.; Sugawara, T.; Iwamura, H. J. Org. Chem. 1981, 46, 4948. (20) Hughes, D. L. Org. React. 1992, 42, 335. (21) Martin, S. F.; Dodge, J. A. Tetrahedron Lett. 1991, 32, 3017. (22) Bartlett, P. A.; Ting, P. C. J. Org. Chem. 1986, 51, 2230.

Stereocontrolled Preparation of Spirocyclic Ethers Scheme 5

J. Org. Chem., Vol. 61, No. 22, 1996 7863

The tricyclic exomethylene ethers 41 and 48 display diagnostic spectral characteristics similar to those of their lower analogues. In both of the examples, the axial R-oxy proton depicted in conformers G and H exhibited unmistakable NOE interactions with the proximal allylic methylene protons. Discussion of Results The stereoelectronic considerations central to a proper assessment of the above results are found in Scheme 1. Conformational analysis of the oxonium ion derived from 16 suggests that minimization of the nonbonded steric interactions would be best realized by projecting the methyl group pseudoequatorially as in I. Intramolecular

with 10 followed by warming to room temperature delivered a 25:1 mixture of 41 and 42. Under the aegis of p-toluenesulfonic acid, the isomerization of 41 to 42 was readily accomplished, to the point where 42 predominated heavily (99:1). In order to avail ourselves of 47, keto ester 4323 was transformed via 44 to the primary alcohol 45, oxidation of which with pyridinium chlorochromate afforded the cis aldehyde 46 (Scheme 6). The conversion of 46 into sulfone 47 ensued. In this example, alkylation and desulfonylative cyclization was met with the exclusive formation of 48. Remarkably, 48 is stable to p-toluenesulfonic acid in CH2Cl2. Evidently, the thermodynamics in the cis series differs significantly from that resident in the trans-fused counterpart. Scheme 6

capture of the allylsilane likely proceeds under kinetic control without a necessary link to product stability. Since antiperiplanar attack, as shown in I, would lead via a chairlike transition state to the product, our expectation was that this pathway would be followed.4,9b,12 Indeed, the conversion to 17 and 18 does indicate that C-C bond formation proceeds exclusively from that π-surface of the oxonium ion anti to the alkyl substituent on the tetrahydropyran ring. In the case of 19, the intervention of conformer J and the parallel stereoelectronic control during cyclization concisely rationalize the production of 20 and 21. Sulfone 22 represents the only example where two stereochemically distinctive pathways were found to operate. The favored formation of 23 suggests that cyclization of oxonium ion K via a chairlike transition state predominates over the second chairlike option L. While this bias is in line with the energetic advantage that should accrue to the equatorial methyl conformer, the dual operation of the two reaction channels implies that a 5-methyl substituent is less controlling of the (23) Stork, G.; Brizzolara, A.; Landesman, H.; Szmuszkovicz, J.; Terrell, R. J. Am. Chem. Soc. 1963, 85, 207.

7864 J. Org. Chem., Vol. 61, No. 22, 1996

conformational energies than one at either C-4 or C-6. A similar dropoff in the stereoselectivity has been noted by Mukaiyama for reductions of structurally related oxonium intermediates with triethylsilane.24 The trans-4,6-dimethyl example 25 is a particularly intriguing case study. The finding that cyclization operates cleanly cis to the 4-Me and trans to the 6-Me suggests that the substitution resident in the associated oxonium ion is reinforcing of stereocontrol rather than being mismatched. However, the configuration set at the spiro carbon can be reconciled only via the chairlike closure depicted in M or the boatlike alternative N. Although we are not in a position to distinguish between these alternatives, attention is called to the potential enthalpic disadvantage associated with the developing 1,3-diaxial interaction in M. This feature is not present in N, which instead must face the energetic costs of proceeding to a twist-boat product. This particular geometry may be more readily accommodated than what was initially assumed since the methylenecyclohexane ring encounters minimal steric compression against either methyl group in that arrangement.12,25 For the trans and cis bicyclic sulfones 40 and 47, analysis of the stereochemical data leads one to conclude that stereoinduction is totally controlled by cyclization on that face of their oxonium ions syn to the angular R-oxy proton. By inspection of G and H, it can easily be gauged that chairlike transition states are adopted in both instances. Although ground state torsional interactions can be complex in their origin, extrapolation of these data to transition structures is often done for estimation purposes. However, there is yet no way to estimate the torsional effects of the developing bonds in the developing transition states. Some of the difficulty lies in our lack of knowledge of the extent of rehybridization, which is dependent in turn on the earliness or lateness of the transition state. Notwithstanding, the present study provides a foundation upon which predictions involving the nucleophilic capture of cyclic oxonium ions may be more soundly based.

Experimental Section General Considerations. Melting points are uncorrected. The column chromatographic separations were performed with Woelm silica gel (230-400 mesh). Solvents were reagent grade and in most cases dried prior to use. The purity of all of the compounds was shown to be >95% by TLC and highfield 1H NMR (300 MHz) and 13C NMR (75 MHz). The highresolution mass spectra were obtained at The Ohio State University Campus Chemical Instrumentation Center. Elemental analyses were performed at the Scandinavian Microanalytical Laboratory, Herlev, Denmark, and at Atlantic Microlab, Inc., Norcross, GA. Alkylation-Cyclization of Dihydropyran. A cold (-78 °C), magnetically stirred solution of dihydropyran (461 mg, 5.48 mmol) in dry THF (30 mL) was treated with tertbutyllithium (3.9 mL of 1.7 M in pentane), stirred for 10 min, warmed to 0 °C for 30 min, and returned to -78 °C, at which point iodide 105 (506 mg, 1.79 mmol) was introduced. The reaction mixture was allowed to warm slowly to rt, stirred for 15 h, and quenched with water (20 mL). The organic layer was separated and the aqueous layer was extracted with ether (24) Homma, K.; Takenoshita, H.; Mukaiyama, T. Bull. Chem. Soc. Jpn. 1990, 63, 1898. (25) (a) Cieplak, P.; Howard, A. E.; Powers, J. P.; Rychnovsky, S. D.; Kollman, P. A. J. Org. Chem. 1996, 61, 3662. (b) Rychnovsky, S. D.; Yang, G.; Powers, J. P. J. Org. Chem. 1993, 58, 5251. (c) Rychnovsky, S. D.; Skalitzky, D. J. Tetrahedron Lett. 1990, 31, 945.

Paquette and Tae (2 × 30 mL). The combined organic solutions were dried over anhydrous K2CO3 and concentrated to leave a residue that was flash chromatographed on basic alumina (elution with 8:1 hexanes/ethyl acetate with 1% triethylamine). The product 11 was isolated as a colorless liquid (178 mg, 65%). A CH2Cl2 solution (15 mL) of 11 (91 mg, 0.382 mmol) was stirred with Dowex-50 (30 mg) for 5 h, passed through a small pad of Celite, and concentrated. Chromatography of the crude product on silica gel (elution with 4:1 hexanes/ethyl acetate) gave 57 mg (90%) of an 18:49:1 mixture of 12, 13, and 14 as a colorless oil. For 12: IR (neat, cm-1) 1660, 1500, 1095, 1060; 1H NMR (300 MHz, CDCl3) δ 4.71 (s, 1 H), 4.66 (s, 1 H), 3.40 (m, 2 H), 2.34 (d, J ) 13.2 Hz, 1 H), 2.17 (d, J ) 13.2 Hz, 1 H), 2.2-2.1 (m, 2 H), 1.85 (m, 1 H), 1.7-1.6 (m, 4 H), 1.6-1.4 (m, 5 H); 13C NMR (75 MHz, CDCl ) ppm 146.2, 109.2, 73.4, 61.0, 44.0, 3 34.6, 34.43, 34.39, 26.0, 22.3, 19.0; MS m/z (M+) calcd 116.1358, obsd 116.1353. Anal. Calcd for C11H18O: C, 79.45; H, 10.92. Found: C, 79.64; H, 10.96. For 13: 1H NMR (300 MHz, CDCl3) distinctive signals appear at δ 5.35 (br s, 1 H), 2.00 (s, 2 H), and 1.63 (s, 3 H); 13C NMR (75 MHz, CDCl3) ppm key absorptions at 131.2, 120.2, 71.4, 60.9, 23.7. For 14: 1H NMR (300 MHz, CDCl3) distinctive signals appear at δ 5.54 (s, 1 H) and 1.68 (s, 3 H); 13C NMR (75 MHz, CDCl3) ppm key absorptions at 137.9, 125.3, 70.6, 61.4, 23.8. General Procedure for the Alkylation-Cyclization of the 2-(Benzene-sulfonyl)tetrahydropyrans. A. A cold (-78 °C), magnetically stirred solution of 1514 (620 mg, 2.74 mmol) in dry THF (10 mL) was treated dropwise with tertbutyllithium (1.9 mL of 1.7 M in hexanes, 3.23 mmol) under N2. After 1 h, iodide 10 (773 mg, 2.74 mmol) was slowly introduced, and the reddish reaction mixture was stirred for 1 h at -78 °C and for 6 h at rt prior to being quenched with water (20 mL). The separated aqueous layer was extracted with ether (2 × 30 mL), and the combined organic layers were dried and carefully evaporated. Column chromatography of the residue on silica gel (elution with 20:1 hexanes/ether) gave 230 mg (51%) of a 64:25:11 mixture of 12, 13, and 14. B. The 6-Methyl Derivative. Comparable treatment of 1615 (890 mg, 3.70 mmol) furnished 362 mg (54%) of a 98:2 mixture of 17 and 18 as a colorless oil. For 17: IR (neat, cm-1) 1655, 1450, 1250, 1080, 1055; 1H NMR (300 MHz, CDCl3) δ 4.68 (s, 1 H), 4.60 (s, 1 H), 3.70 (ddt, J ) 11.4, 2.2, 6.1 Hz, 1 H), 2.37 (s, 2 H), 2.11 (t, J ) 6.2 Hz, 2 H), 1.8-1.4 (m, 10 H), 1.10 (d, J ) 6.1 Hz, 3 H); 13C NMR (75 MHz, CDCl3) ppm 146.7, 108.8, 74.1, 65.4, 40.3, 39.9, 34.7, 33.6 (2 C), 32.7, 22.7, 19.3; MS m/z (M+) calcd 180.1514, obsd 180.1504. Anal. Calcd for C12H20O: C, 79.94; H, 11.18. Found: C, 79.84; H, 11.13. For 18: 1H NMR (300 MHz, CDCl3) distinctive signals appear at δ 5.33 (s, 1 H), 2.07 (s, 2 H), and 1.61 (s, 3 H); 13C NMR (75 MHz, CDCl3) ppm key absorptions at 131.4 and 120.9. C. The 4-Methyl Derivative. Analogous processing of 1916 (247 g, 1.03 mmol) provided 104 mg (56%) of an 81:19 mixture of 20 and 21 as a colorless oil. For 20: IR (neat, cm-1) 1652, 1445, 1374, 1325, 1162, 1086; 1H NMR (300 MHz, C D ) δ 4.79 (s, 1 H), 4.70 (s, 1 H), 3.63 6 6 (ddd, J ) 1.6, 5.1, 11.8 Hz, 1 H), 3.48 (dt, J ) 2.5, 12.2 Hz, 1 H), 2.46 (d, J ) 13.3 Hz, 1 H), 2.03 (d, J ) 13.3 Hz, 1 H), 2.201.80 (m, 3 H), 1.67 (m, 1 H), 1.60-1.35 (m, 4 H), 1.23 (m, 1 H), 1.05 (m, 1 H), 0.80 (m, 1 H), 0.76 (d, J ) 6.2 Hz, 3 H); 13C NMR (75 MHz, CDCl3) ppm 146.3, 109.2, 74.0, 60.7, 42.9, 39.9, 39.6, 34.62, 34.57, 25.1, 22.7, 22.5; MS m/z (M+) calcd 180.1514, obsd 180.1511. Anal. Calcd for C12H20O: 79.94; H, 11.18. Found: C, 79.66; H, 11.16. For 21: 1H NMR (300 MHz, CDCl3) distinctive signals appear at δ 5.38 (s, 1 H), 2.14 (s, 2 H), and 1.66 (s, 3 H); 13C NMR (75 MHz, CDCl3) ppm one of the olefnic carbons is evident at 120.3. D. The 5-Methyl Derivative. From the alkylation of 22 (927 mg, 3.86 mmol), there was obtained 368 mg (53%) of a 3.7:1 mixture of 23 and 24 as a colorless oil: IR (neat, cm-1) 1652, 1439, 1356, 1084; 1H NMR (300 MHz, CDCl3) δ 4.71 (s, 1 H), 4.64 (s, 1 H), 3.56 (m, 1 H), 3.24 (dd, J ) 10.5, 11.5 Hz,

Stereocontrolled Preparation of Spirocyclic Ethers 0.78 H, major), 3.15 (dd, J ) 10.3, 11.5 Hz, 0.22 H, minor), 2.42 (d, J ) 13.3 Hz, 1 H), 2.23 (d, J ) 12.9 Hz, 1 H), 2.201.95 (m, 2 H), 1.75-1.15 (series of m, 9 H), 0.82 (d, J ) 6.5 Hz, 3 H); 13C NMR (75 MHz, CDCl3) (for 23) ppm 146.2, 109.2, 73.1, 67.1, 39.5, 38.9, 34.6, 33.5, 30.8, 27.5, 22.6, 17.3; 13C NMR (75 MHz, CDCl3) (for 24) ppm 146.2, 109.1, 73.1, 67.4, 48.1, 38.9, 34.3, 30.6, 29.4, 27.8, 22.2, 17.3; MS m/z (M+) calcd 180.1514, obsd 180.1506. Anal. Calcd for C12H20O: C, 79.94; H, 11.18. Found: C, 79.82; H, 11.10. E. The trans-4,6-Dimethyl Derivative. Comparable treatment of 25 (931 mg, 3.66 mmol) gave rise to 625 mg (88%) of a 7.4:1 mixture of 26 and 27 as a colorless oil. For 26: IR (neat, cm-1) 1650, 1450, 1380, 1075; 1H NMR (300 MHz, CDCl3) δ 4.68 (s, 1 H), 4.61 (s, 1 H), 4.00-3.65 (m, 1 H), 2.29 (d, J ) 13.0 Hz, 2 H), 2.30-1.80 (series of m, 3 H), 1.80-1.35 (series of m, 6 H), 1.35-1.10 (m, 2 H), 1.15 (d, J ) 6.3 Hz, 3 H), 1.01 (d, J ) 7.0 Hz, 3 H); 13C (75 MHz, CDCl3) ppm 146.8, 108.9, 74.6, 63.1, 45.9, 39.6, 39.5, 38.3, 34.5, 23.6, 23.1, 22.6, 21.8; MS m/z (M+) calcd 194.1670, obsd 194.1656. Anal. Calcd for C13H22O: C, 80.35; H, 11.41. Found: C, 80.20; H, 11.47. For 27: 1H NMR (300 MHz, CDCl3) distinctive signals seen at δ 5.40 (m, 1 H), 2.08 (d, J ) 13.0 Hz, 2 H), and 1.63 (m, 3 H); 13C NMR (75 MHz, CDCl3) ppm key absorptions at 131.7, 120.1, 72.3, 63.8. F. Starting from 40. Analogous processing of 40 (360 mg, 1.28 mmol) afforded 220 mg (78%) of a 25:1 mixture of 41 and 42. For 41: colorless oil; IR (neat, cm-1) 1651, 1449, 1365, 1105, 1074; 1H NMR (300 MHz, CDCl3) δ 4.68 (s, 1 H), 4.61 (s, 1 H), 3.20 (m, 1 H), 2.45 (d, J ) 13.0 Hz, 1 H), 2.31 (d, J ) 13.0 Hz, 1 H), 2.15 (m, 2 H), 1.80-0.0 (series of m, 17 H); 13C NMR (75 MHz, CDCl3) ppm 148.8, 109.0, 74.2, 73.9, 42.5, 40.3, 40.2, 34.8, 33.6, 33.2, 31.8, 26.6, 26.0, 25.3, 22.8; MS m/z (M+) calcd 220.1827, obsd 220.1823. Anal. Calcd for C15H24O: C, 81.76; H, 10.98. Found: C, 81.89; H, 11.03. A 50 mg (0.23 mmol) sample of 41 was stirred in CH2Cl2 (20 mL) containing a crystal of p-toluenesulfonic for 15 h resulting in isomerization to a 1:99 mixture of 41 and 42. For 42: colorless oil; IR (neat, cm-1) 1680, 1445, 1355, 1100, 1085; 1H NMR (300 MHz, CDCl3) δ 5.33 (s, 1 H), 3.20 (m, 1 H), 2.33 (d, J ) 17.0 Hz, 1 H), 2.15-1.90 (m, 3 H), 1.85-1.40 (m, 6 H), 1.66 (s, 3 H), 1.40-0.85 (series of m, 9 H); 13C NMR (75 MHz, CDCl3) ppm 131.8, 120.2, 74.3, 72.5, 42.5, 37.1, 36.4, 33.3, 32.8, 31.8, 26.9, 26.0, 25.3, 23.9, 23.2; MS m/z (M+) calcd 220.1827, obsd 220.1828. G. Starting from 47. Comparable treatment of 47 (252 mg, 0.90 mmol) yielded 168 mg (85%) of 48 as a colorless oil: IR (neat, cm-1) 1650, 1445, 1079, 1060, 1051, 1014; 1H NMR (300 MHz, CDCl3) δ 4.67 (s, 1 H), 4.59 (s, 1 H), 3.82 (m, 1 H), 2.48 (d, J ) 12.9 Hz, 1 H), 2.26 (d, J ) 12.9 Hz, 1 H), 2.10 (m, 2 H), 1.90-1.10 (series of m, 17 H); 13C NMR (75 MHz, CDCl3) ppm 147.0, 108.8, 74.1, 66.4, 40.4, 40.1, 34.8, 34.7, 32.3, 27.3, 26.2, 25.2, 24.9, 22.7, 20.3; MS m/z (M+) calcd 220.1827, obsd 220.1828. Anal. Calcd for C15H24O: C, 81.76; H, 10.98. Found: C, 81.49; H, 10.79. Tetrahydro-2-(benzenesulfonyl)-5-methyl-2H-pyran (22). A cold (0 °C) solution of acrolein (6.68 mL, 0.10 mol) in dry benzene (120 mL) was treated with trimethylsilyl bromide (15.8 mL, 0.12 mol), stirred for 30 min at 0 °C, allowed to warm to rt, and stirred for an additional hour prior to the addition of allyl alcohol (20 mL, 0.30 mol) and p-toluenesulfonic acid (500 mg). The mixture was refluxed under a Dean-Stark trap for 15 h, cooled, and evaporated. Chromatography of the residue on silica gel (elution with 4:1 hexanes/ethyl acetate) gave 28 as a colorless liquid (20.8 g, 88%): IR (neat, cm-1) 1647, 1423, 1388, 1346, 1263, 1215, 1117, 1045; 1H NMR (300 MHz, CDCl3) δ 5.90 (m, 2 H), 5.25 (m, 4 H), 4.77 (t, J ) 5.5 Hz, 1 H), 4.10 (m, 4 H), 3.43 (t, J ) 6.8 Hz, 2 H), 2.18 (dt, J ) 5.6, 6.7 Hz, 2 H); 13C NMR (75 MHz, CDCl3) ppm 134.3, 116.9, 100.5, 67.2, 36.8, 28.5; MS m/z (M+) calcd 234.0255, obsd 234.0271. A refluxing benzene solution (100 mL) of 28 (20.8 g, 88.5 mmol) and AIBN (100 mg) was treated dropwise with tri-nbutyltin hydride (28.5 mL, 106 mmol) in benzene (22 mL) over 4 h. After an additional hour of heating, the reaction mixture

J. Org. Chem., Vol. 61, No. 22, 1996 7865 was concentrated, and the residue was filtered through silica gel (elution with 10:1 hexanes/ethyl acetate) to give 29 which was stirred overnight at rt with benzenesulfinic acid (20.0 g, 107 mmol) and CaCl2 (30.0 g) in CH2Cl2 (200 mL). The reaction mixture was filtered and evaporated, and the product was purified by chromatography on silica gel (elution with 8:1 hexanes/ethyl acetate) to give 22 as a colorless solid, mp 103105 °C (10.8 g, 53% for 2 steps). IR (neat, cm-1) 1447, 1307, 1148, 1081, 1022; 1H NMR (300 MHz, CDCl3) δ 7.87 (m, 2 H), 7.65 (m, 1 H), 7.52 (m, 2 H), 4.28 (br d, J ) 11.5 Hz, 1 H), 3.91 (br d, J ) 11.3 Hz, 1 H), 2.95 (t, J ) 11.1 Hz, 1 H), 2.15 (m, 1 H), 1.95 (m, 1 H), 1.65 (m, 2 H), 1.15 (m, 1 H), 0.73 (d, J ) 6.7 Hz, 3 H); 13C NMR (75 MHz, CDCl3) ppm 136.2, 133.8, 129.4, 128.7, 91.3, 75.0, 30.4, 29.7, 23.8, 16.4; MS m/z (M+ - SO2Ph) calcd 99.0809, obsd 99.0807. Anal. Calcd for C12H13O3S: C, 59.98; H, 6.71. Found: C, 60.06; H, 6.68. (1R*,3S*)-5-(tert-Butyldimethylsiloxy)-1,3-dimethylpentyl p-Nitrobenzoate (33). A solution of 30 (8.56 g, 66.8 mmol) in ether (200 mL) was stirred at 0 °C with lithium aluminum hydride (2.53 g, 66.8 mmol) for 1.5 h, quenched with saturated Na2SO4 solution at 0 °C, filtered, and concentrated. Chromatography of the residue on silica gel (elution with 1:1 hexanes/ethyl acetate) gave 31 as a colorless oil. The diol was dissolved in CH2Cl2 (150 mL) containing triethylamine (14.0 mL, 100 mmol), cooled to 0 °C, and treated with a solution of tert-butyldimethylsilyl chloride (10.1 g, 66.8 mmol) and 4-(dimethylamino)pyridine (50 mg) in CH2Cl2 (50 mL). After stirring overnight at rt, the reaction mixture was washed with 2 N HCl (100 mL) and brine (50 mL), dried, and concentrated. Chromatography of the residue on silica gel (elution with 4:1 hexanes/ethyl acetate) afforded 11.8 g (72%) of 32 as a colorless oil: IR (neat, cm-1) 3360, 1463, 1378, 1255, 1096, 1005, 901; 1H NMR (300 MHz, CDCl ) δ 3.88 (m, 1 H), 3.65 (m, 2 H), 3 1.72 (s, 1 H), 1.85-1.65 (m, 1 H), 1.60-1.30 (m, 3 H), 1.251.05 (m, 1 H), 1.17 (d, J ) 6.1 Hz, 3 H), 0.91 (d, J ) 6.8 Hz, 3 H), 0.88 (s, 9 H), 0.04 (s, 6 H); 13C NMR (75 MHz, CDCl3) ppm 66.0, 61.3, 46.9, 40.3, 26.6, 26.0, 24.3, 19.9, 18.4, -5.2; MS (M+ - OH) calcd 229.1988, obsd 229.1972. To a mixture of triphenylphosphine (15.1 g, 57.5 mmol) and p-nitrobenzoic acid (9.61 g, 57.5 mmol) in THF (200 mL) was added 32 (11.8 g, 47.9 mmol) dissolved in THF (30 mL), and subsequently after cooling to -30 °C, a THF solution (70 mL) of diethyl azodicarboxylate (9.19 mL, 57.5 mmol) was added. The reaction mixture was warmed slowly to 0 °C and after 2.5 h was quenched with saturated NaHCO3 solution (150 mL). The separated aqueous layer was extracted with ether (2 × 100 mL), and the combined organic solutions were dried and evaporated. Purification of the product by initial addition of 1:3 ether/hexanes to precipitate the triphenylphosphine oxide and subsequent chromatography on silica gel (elution with 8:1 hexanes/ethyl acetate) furnished 16.6 g (88%) of 33, a colorless oil: IR (neat, cm-1) 1723, 1608, 1471, 1349, 1276, 1101, 1015; 1H NMR (300 MHz, CDCl ) δ 8.25 (m, 4 H), 5.29 (m, 1 H), 3 3.62 (m, 2 H), 1.80-1.55 (m, 4 H), 1.35 (m, 1 H), 1.35 (d, J ) 6.2 Hz, 3 H), 0.95 (d, J ) 6.4 Hz, 3 H), 0.83 (s, 9 H), 0.00 (s, 6 H); 13C NMR (75 MHz, CDCl3) ppm 164.1, 150.4, 136.2, 130.6, 123.4, 71.6, 60.7, 43.2, 39.6, 26.4, 25.8, 20.1, 19.7, 18.2, -5.4; MS m/z (M+) calcd 395.2128, obsd 395.2112. Anal. Calcd for C20H33NO5Si: C, 60.73; H, 8.41. Found: C, 60.88; H, 8.47. Tetrahydro-2-(benzenesulfonyl)-trans-4,5-dimethyl2H-pyran (25). A solution of 33 (16.6 g, 42.0 mmol) in ether (200 mL) was treated with lithium aluminum hydride (3.20 g, 84.0 mmol) at 0 °C. After 1.5 h, saturated Na2SO4 solution was introduced with ice cooling, and the reaction mixture was filtered and concentrated. Chromatography of the residue on silica gel (elution with 4:1 hexanes/ethyl acetate) gave 34 (8.63 g, 83%) as a colorless oil: IR (neat, cm-1) 3360, 1472, 1378, 1255, 1097, 1006, 981; 1H NMR (300 MHz, CDCl3) δ 3.91 (m, 1 H), 3.65 (m, 2 H), 1.74 (s, 1 H), 1.80-1.55 (m, 2 H), 1.401.25 (m, 3 H), 1.17 (d, J ) 6.1 Hz, 3 H), 0.91 (d, J ) 6.6 Hz, 3 H), 0.89 (s, 9 H), 0.05 (s, 6 H); 13C NMR (75 MHz, CDCl3) ppm 65.9, 61.3, 46.8, 39.4, 26.6, 25.9, 23.6, 20.5, 18.3, -5.3; MS m/z (M+ - OH) calcd 229.1988, obsd 229.1993. A mixture of 34 (8.63 g, 35.0 mmol) and p-toluenesulfonic acid (50 mg) in CH2Cl2 (100 mL) was treated with dihydropyran (6.4 mL, 70 mmol), stirred for 2 h, concentrated in vacuo,

7866 J. Org. Chem., Vol. 61, No. 22, 1996 and subjected directly to chromatography on silica gel (elution with 10:1 hexanes/ethyl acetate). The resulting 35 was dissolved in THF (150 mL) and treated with 4 Å molecular sieves (20 g) and tetra-n-butylammonium fluoride hydrate (13.2 g, 50.4 mmol). After 3 h of stirring at rt, the reaction mixture was decanted, washed with water (100 mL) and brine (50 mL), dried, and concentrated. Column chromatography of the residue on silica gel (elution with 2:1 hexanes/ethyl acetate) gave 36 as a colorless oil (6.00 g, 78%): IR (neat, cm-1) 3417, 1454, 1378, 1260, 1200, 1120, 1076, 1023; 1H NMR (300 MHz, C6D6) δ 4.71 (m, 0.6 H), 4.58 (m, 0.4 H), 4.00-3.75 (m, 2 H), 3.75-3.45 (m, 2 H), 3.45-3.25 (m, 1 H), 2.80 (br s, 1 H), 1.95 (m, 1 H), 1.85-1.00 (m, 10 H), 1.27 (d, J ) 6.2 Hz, 1.8 H), 1.01 (d, J ) 6.0 Hz, 1.2 H), 0.90 (d, J ) 6.7 Hz, 1.2 H), 0.85 (d, J ) 6.6 Hz, 1.8 H); 13C NMR (75 MHz, C6D6) ppm 98.6, 96.6, 72.7, 69.6, 63.0, 62.1, 60.6, 60.4, 45.5, 44.9, 40.4, 40.0, 31.7, 31.6, 26.7, 26.3, 26.0, 25.8, 22.2, 20.8, 20.4, 20.3, 19.9, 19.7; MS m/z (M+) calcd 216.1725, obsd 216.1706. A mixture of 36 (2.88 g, 13.3 mmol), 4 Å molecular sieves (6.0 g), and sodium acetate (100 mg) in CH2Cl2 (100 mL) was treated with pyridinium chlorochromate (4.30 g, 20.0 mmol), stirred for 1.5 h, loaded directly atop a column of silica gel, and eluted with 10:1 hexanes/ethyl acetate to give the aldehyde as a colorless oil. This material was stirred overnight with benzenesulfinic acid (5.00 g, 26.9 mmol) and calcium chloride (10.0 g) in CH2Cl2 (150 mL). The reaction mixture was suctionfiltered, the filtrate was evaporated, and the residue was purified by chromatography on silica gel (elution with 8:1 hexanes/ethyl acetate) to provide 2.53 g (75%) of 25 as a white solid, mp 67-69 °C: IR (neat, cm-1) 1446, 1383, 1308, 1150, 1074; 1H NMR (300 MHz, CDCl3) δ 7.90 (m, 2 H), 7.65 (m, 1 H), 7.55 (m, 2 H), 4.62 (dd, J ) 3.4, 11.5 Hz, 0.4 H, minor), 4.57 (dd, J ) 3.1, 11.4 Hz, 0.6 H, major), 4.30 (m, 0.4 H, minor), 3.60 (dqq, J ) 2.6, 6.1, 11.8 Hz, 0.6 H, major), 2.25 (m, 0.4 H, minor), 2.10 (m, 0.6 H, major), 1.95-1.70 (m, 1.4 H), 1.501.25 (m, 2.6 H), 1.08 (d, J ) 6.2 Hz, 2.4 H), 1.06 (d, J ) 6.3 Hz, 1.8 H), 0.96 (d, J ) 6.5 Hz, 1.8 H); 13C NMR (75 MHz, CDCl3) ppm (major) 136.2, 133.6, 129.2, 128.5, 88.0, 69.8, 37.7, 28.6, 23.7, 21.4, 17.5; (minor) 136.4, 133.6, 129.6, 128.7, 86.8, 70.1, 37.4, 30.3, 25.0, 22.0, 17.9; MS m/z (M+ - SO2C6H5) calcd 113.0966, obsd 113.0945. (4aR*,8aS*)-Hexahydro-2-(phenylsulfonyl)chroman (40). Into a methanol solution (100 mL) of 37 (1.50 g, 9.72 mmol) was passed ozone for 1.5 h at -78 °C. The reaction mixture was treated with dimethyl sulfide (2.00 mL, 27.2 mmol), allowed to warm to rt overnight, and freed of solvent. The residue was subjected to chromatography on silica gel (elution with 4:1 hexanes/ethyl acetate) and afforded a 1:3 mixture of 38 and 39 which was immediately dissolved in CH2Cl2 (100 mL) and stirred overnight with benzenesulfinic acid (3.00 g, 16.1 mmol) and CaCl2 (6.00 g). Following the predescribed workup, 1.75 g (64%) of 40 was isolated as a colorless liquid: IR (neat, cm-1) 1450, 1320, 1155, 1100, 1090; 1H NMR (300 MHz, C6D6) δ 8.01 (m, 2 H), 7.00 (m, 3 H), 4.17 (dd, J ) 2.4, 11.7 Hz, 1 H), 2.35 (m, 1 H), 2.10 (m, 1 H), 1.7-0.4 (series of m, 12 H); 13C NMR (75 MHz, C6D6) ppm 137.9, 133.3, 130.1, 128.6, 91.9, 82.9, 40.6, 32.2, 31.1, 29.3, 25.6, 24.8, 24.7; MS m/z (M+ - SO2C6H5) calcd 139.1122, obsd 139.1121. (4aR*,8aR*)-Hexahydro-2-(phenylsulfonyl)chroman (47). A solution of 43 (21.3 g, 0.107 mmol) in anhydrous ether (200 mL) was treated with lithium aluminum hydride (8.15

Paquette and Tae g, 0.122 mol) at 0 °C, stirred for 1.5 h, quenched with saturated Na2SO4 solution at 0 °C, filtered, and concentrated. Chromatography of the product gave the diol, which was dissolved in CH2Cl2 (250 mL) containing triethylamine (22.0 mL, 0.161 mol), and treated with a solution of tert-butyldimethylsilyl chloride (16.0 g, 0.107 mmol) and 4-(dimethylamino)pyridine (50 mg) at 0 °C. After being stirred overnight, the reaction mixture was washed with 2 N HCl (100 mL) and brine (50 mL) and then dried and concentrated. Column chromatography of the residue on silica gel (elution with 4:1 hexanes/ethyl acetate) gave 12.0 g (41%) of 44 as a colorless oil: IR (neat, cm-1) 3405, 1471, 1388, 1255, 1102, 974; 1H NMR (300 MHz, CDCl3) δ 3.87 (m, 1 H), 3.60 (t, J ) 6.5 Hz, 2 H), 1.85-1.15 (series of m, 14 H), 0.90 (s, 3 H), 0.88 (s, 6 H), 0.08 (s, 2 H), 0.03 (s, 4 H); 13C NMR (75 MHz, CDCl3) ppm 69.2, 63.5, 41.1, 33.0, 30.2, 28.0, 26.8, 25.9, 25.6, 25.2, 20.5, 18.3, -3.6, -5.3; MS m/z (M+) calcd 272.2172, obsd 272.2179. A solution of 44 (2.30 g, 8.44 mmol) and p-toluenesulfonic acid (50 mg) in CH2Cl2 (50 mL) was treated with dihydropyran (1.20 mL, 13.2 mmol), stirred for 2 h, and concentrated in vacuo. Chromatography of the residue on silica gel (elution with 10:1 hexanes/ethyl acetate) gave the diprotected diol as a colorless oil, which was directly dissolved in THF (100 mL) and treated with 4 Å molecular sieves (10 g) and tetra-nbutylammonium fluoride hydrate (4.40 g, 16.8 mmol). After 3 h of stirring at rt, the decanted reaction mixture was washed with water (100 mL) and brine (50 mL), dried, and concentrated. Purification of the residue by chromatography on silica gel (elution with 4:1 hexanes/ethyl acetate) gave 1.15 g (56%) of 45 as a colorless oil: IR (neat, cm-1) 3388, 1453, 1354, 1260, 1199, 1076, 1021, 999; 1H NMR (300 MHz, C6D6) δ 4.66 (t, J ) 3.5 Hz, 0.45 H), 4.58 (t, J ) 3.9 Hz, 0.55 H), 4.00-3.30 (m, 5 H), 2.30-2.00 (series of m, 20 H); 13C NMR (75 MHz, C6D6) ppm 100.2, 95.9, 77.5, 72.5, 62.9 (2 C), 62.3 (2 C), 41.4, 41.3, 32.3, 31.8, 31.5, 30.9, 30.8, 28.8, 28.3, 28.0, 27.9, 26.1, 25.9, 25.7, 25.3, 22.2, 21.2, 20.5, 20.4, 20.1; MS m/z (M+) calcd 242.1882, obsd 242.1882. Oxidation of 45 (1.15 g, 4.75 mmol) with pyridinium chlorochromate (1.53 g, 7.10 mmol) as before and exposure of the resulting aldehyde to benzenesulfinic acid (3.00 g, 16.1 mmol) and CaCl2 (6.0 g) in the customary manner furnished 47 (878 mg, 66%) as a colorless oil: IR (neat, cm-1) 1446, 1310, 1307, 1150, 1060; 1H NMR (300 MHz, CDCl3) δ 8.05 (m, 2 H), 6.95 (m, 3 H), 4.19 (dd, J ) 3.3, 11.4 Hz, 1 H), 3.00 (m, 1 H), 1.85-0.75 (series of m, 13 H); 13C NMR (75 MHz, CDCl3) ppm 137.4, 133.3, 130.2, 128.5, 92.4, 76.7, 33.9, 31.6, 28.0, 25.9, 24.9, 20.5, 19.4; MS m/z (M+ - C6H5SO2) calcd 139.1122, obsd 139.1134.

Acknowledgment. We thank the National Science Foundation and Hoechst Marion Roussel for financial support. Supporting Information Available: Copies of the high field 1H and 13C NMR spectra for those compounds lacking combustion data (20 pages). This material is contained in libraries on microfiche, immediately follows this article in the microfilm version of the journal, and can be ordered from the ACS; see any current masthead page for ordering information. JO961306K